Big Chemical Encyclopedia

Chemical substances, components, reactions, process design ...

Articles Figures Tables About

A simplified hamiltonian

In some cases we can simplify the above expression. Thus, by using that we often have ip Cg Spg, where p,q = P, (f , p, from eqs.(7.35)-(7.40) we obtain the following approximations  [Pg.161]

The reaction volume hamiltonian can be used for large systems in which the reaction is a three center reaction. For such a system it will be advantageous to treat the SN — 9 perpendicular vibrations within the second quantization approach. By minimizing the gradient we define a reference volume in which all reaction following motions as e.g. umbrella and other concerted motion is taken care of through the reference position p). Small am- [Pg.162]

With the RP method we may treat reactions involving many atoms and the method is the from a computational point of view the most economic way of doing so, the reason being that the information needed from electronic structure calculations is just gradients and hessians along the path. Thus we may avoid the construction of the full PES. The method offers dynamical [Pg.162]

However, the restrictions on the dynamics posed by the inclusion of only quadratic terms in the expansion of the perpendicular vibrational motions, the fact that the reaction path itself may be poorly deflned in certain regions, bifurcations of the path should not be overlooked. In such cases it is possible to switch to the reaction surface or the reaction volume methods. These methods are more demanding from the point of view of the potential energy surface and also as far as the dynamical requirements are concerned. But they nevertheless reduce the problem considerably by demanding only the full information in a 2 or 3D space and the harmonic displacement of the spectator atoms connected to the reaction center. [Pg.163]

In order to include the 9 constraints we follow the prescription given in [10]. Thus introducing the expression (7.30) the constraints will be [Pg.163]


This is a simplified Hamiltonian that ignores the direct interaction of any nuclear spins with the applied field, B. Because of the larger coupling, Ah to most transition metal nuclei, however, it is often necessary to use second-order perturbation theory to accurately determine the isotropic parameters g and A. Consider, for example, the ESR spectrum of vanadium(iv) in acidic aqueous solution (Figure 3.1), where the species is [V0(H20)5]2+. [Pg.44]

Spin-Orbit Coupling For the derivation of selection rules, it is sufficient to employ a simplified Hamiltonian. To this end, we rewrite each term in the microscopic spin-orbit Hamiltonians in form of a scalar product between an appropriately chosen spatial angular momentum 2 and a spin angular momentum S... [Pg.146]

As a first example we consider a nondegenerate single-channel system with a simplified Hamiltonian in which continuum-continuum interactions have been neglected ... [Pg.492]

The coefficients are combined with the radial term to construct a simplified Hamiltonian describing the d-orbital energies. Note that represents a radial correction due to LFSE only, and that the primary contribution to the radial dependence of the ligand-metal interaction energy comes from the standard... [Pg.29]

Since in EPR only transitions between the lowest levels arc studied, it seems to be a good approximation to treat them as isolated levels. Starting from this assumption. Abragam and Pryce have shown [56, 60] that a simplified Hamiltonian... [Pg.976]

Semiempirical calculations generally use a simplified Hamiltonian operator and incorporate experimental data or a set of parameters that can be adjusted to fit experimental data. [Pg.171]

Let US consider the simplified Hamiltonian in which the nuclear kinetic energy term is neglected. This also implies that the nuclei are fixed at a certain configuration, and the Hamiltonian describes only the electronic degrees of freedom. This electronic Hamiltonian is... [Pg.402]

As a naive or zero-order approximation, we can simply ignore the V12 term and allow the simplified Hamiltonian to operate on the Is orbital of the H atom. The result is... [Pg.236]

In the DC-biased structures considered here, the dynamics are dominated by electronic states in the conduction band [1]. A simplified version of the theory assumes that the excitation occurs only at zone center. This reduces the problem to an n-level system (where n is approximately equal to the number of wells in the structure), which can be solved using conventional first-order perturbation theory and wave-packet methods. A more advanced version of the theory includes all of the hole states and electron states subsumed by the bandwidth of the excitation laser, as well as the perpendicular k states. In this case, a density-matrix picture must be used, which requires a solution of the time-dependent Liouville equation. Substituting the Hamiltonian into the Liouville equation leads to a modified version of the optical Bloch equations [13,15]. These equations can be solved readily, if the k states are not coupled (i.e., in the absence of Coulomb interactions). [Pg.251]

The total Hamiltonian is the sum of the two terms H = H + //osc- The way in which the rate constant is obtained from this Hamiltonian depends on whether the reaction is adiabatic or nonadiabatic, concepts that are explained in Fig. 2.2, which shows a simplified, one-dimensional potential energy surface for the reaction. In the absence of an electronic interaction between the reactant and the metal (i.e., all Vk = 0), there are two parabolic surfaces one for the initial state labeled A, and one for the final state B. In the presence of an electronic interaction, the two surfaces split at their intersection point. When a thermal fluctuation takes the system to the intersection, electron transfer can occur in this case, the system follows the path... [Pg.35]

Semi-empirical methods, such as those implemented in the MOPAC [9] program, simplify the equations considerably by neglecting many terms, but then compensate for this by parameterising some of them so that the calculations reproduce experimental information on, for example, the heat of formation. Once the various approximations are made, the molecular properties to which the parameters are fitted, and the molecules used in the fitting, define a model Hamiltonian, of which the most commonly used are the AMI and the PM3 Hamiltonians found in MOPAC. A major advantage of semi-empirical methods is... [Pg.688]

Abernathy and Sharp employed a similar idea, although in a more simplified form 130). They also worked in terms of a spin Hamiltonian varying with time in discrete steps and let the Hamiltonian contain the Zeeman and the ZFS interactions. They assumed, however, that the ZFS interaction was constant in the molecule-fixed (P) frame and that variation of the Hamiltonian originated only from fluctuation of the P frame with respect to the laboratory frame. These fluctuations were described in terms of Brownian reorientational motion, characterized by a time interval, x, (related to the rotational correlation time x ) and a Gaussian distribution of angular steps. [Pg.84]

In principle, it should be possible to obtain the electronic energy levels of the molecules as a solution of the Schrodinger equation, if inter-electronic and internuclear cross-coulombic terms are included in the potential energy for the Hamiltonian. But the equation can be solved only if it can be broken up into equations which are functions of one variable at a time. A simplifying feature is that because of the much larger mass of the nucleus the motion of the electrons can be treated as independent of that of the nucleus. This is known as the Bom-Oppen-heimer approximation. Even with this simplification, the exact solution has been possible for the simplest of molecules, that is, the hydrogen molecule ion, H + only, and with some approximations for the H2 molecule. [Pg.29]

The theoretical difficulty of making this separation derives from the indistinguish-ability of electrons and the requirement that the total wavefunction be antisymmetric with respect to permutations of the electronic co-ordinates. One approach has been to abandon a full quantum mechanical description in favour of a simplified model hamiltonian which can be conveniently parameterized in terms of experimental quantities. This is the rationale behind Huckel theory, CNDO, and other more sophisticated methods such as MINDO. These techniques have been well documented and reviewed elsewhere (Dewar,1 Pariser, Parr, and Pople,2 Murrell and Hargett,3 etc.) and will not be pursued further here. [Pg.100]

As an approximation to equation (11) it is useful to consider a valence hamiltonian for which only the one-electron operators are transformed and the two-electron interactions remain unchanged. The hamiltonian now simplifies to ... [Pg.103]

In their early study of the ruby spectrum Sugano and Tanabe [44] attributed the zfs to the trigonal anisotropy of the spin-orbit coupling. In a simplified form the anisotropic s.o.c. hamiltonian may be written as ... [Pg.48]

The formal theory of resonances due to Feshbach begins with the decomposition of the Hamiltonian in terms of a projection operator Q [8]. He defines Q as the projection onto the closed-channel space, just like the example of H discussed around Eqs. (4) and (5). Then, QBSs described well by the eigenfunctions Q4> of Eq. (5) with his Q may be called Feshbach resonances." A simplified picture would be that eigenstates Q are supported by some attractive effective potential approaching asymptotically the threshold energy of a closed channel. If this is the case, then the energies EQ of... [Pg.170]

The resultant effective Hamiltonian may be given in a simplified form. First the 3-site permutations simplify on spin space... [Pg.67]

As in the bosonic case, a simplified algorithm can be derived for time-independent system Hamiltonians. In eigenstate representation with eigenstates /i), v) and transition frequencies u) the matrix elements of Axx> t) are given by... [Pg.354]

As Dewar points out in ref. [30a], this derivation is not really satisfactory. A rigorous approach is a simplified version of the derivation of the Hartree-Fock equations (Chapter 5, Section 5.2.3). It starts with the total molecular wavefunction expressed as a determinant, writes the energy in terms of this wavefunction and the Hamiltonian and finds the condition for minimum energy subject to the molecular orbitals being orthonormal (cf. orthogonal matrices, Section 4.3.3). The procedure is explained in some detail in Chapter 5, Section 5.2.3)... [Pg.170]

For abundant nuclei with spin V2, the spectrum is often dominated by heteronuclear or homonuclear dipolar interactions, i.e., the interactions between the magnetic moments of two neighbouring spins. In this case there is no isotropic contribution and q is zero, so that Equation 14.1 simplifies correspondingly. For a two-spin system one obtains a spin Hamiltonian of the form ... [Pg.521]

This is the simplest possible mechanistic model of the PES, derived from an approximate treatment of energy according to eq. (3.69). The FA type of treatment implies that the geminal amplitude-related ES V eqs. (2.78) and (2.81) are fixed at their invariant values eq. (3.7). This corresponds clearly to a simplified situation where all bonds are single ones. Within such a picture, the dependence of the energy on the interatomic distance reduces to that of the matrix elements of the underlying QM (MINDO/3 or NDDO) semiempirical Hamiltonian. [Pg.231]

Whereas the quantum-mechanical molecular Hamiltonian is indeed spherically symmetrical, a simplified virial theorem should apply at the molecular level. However, when applied under the Born-Oppenheimer approximation, which assumes a rigid non-spherical nuclear framework, the virial theorem has no validity at all. No amount of correction factors can overcome this problem. All efforts to analyze the stability of classically structured molecules in terms of cleverly modified virial schemes are a waste of time. This stipulation embraces the bulk of modern bonding theories. [Pg.117]


See other pages where A simplified hamiltonian is mentioned: [Pg.32]    [Pg.69]    [Pg.28]    [Pg.200]    [Pg.150]    [Pg.196]    [Pg.41]    [Pg.161]    [Pg.26]    [Pg.32]    [Pg.69]    [Pg.28]    [Pg.200]    [Pg.150]    [Pg.196]    [Pg.41]    [Pg.161]    [Pg.26]    [Pg.40]    [Pg.41]    [Pg.2297]    [Pg.508]    [Pg.123]    [Pg.110]    [Pg.616]    [Pg.12]    [Pg.57]    [Pg.92]    [Pg.208]    [Pg.163]    [Pg.288]    [Pg.566]   


SEARCH



Simplified

Simplified Hamiltonian

Simplify

© 2024 chempedia.info